You are on page 1of 12

Journal of Loss Prevention in the Process Industries 85 (2023) 105164

Contents lists available at ScienceDirect

Journal of Loss Prevention in the Process Industries


journal homepage: www.elsevier.com/locate/jlp

CFD simulations of vapour cloud explosions using PDRFoam


Manish Dhiman a,c ,1 , Anand Zambare a ,1 , Pratap Sathiah b , Vagesh D. Narasimhamurthy a ,∗
a
Department of Applied Mechanics, Indian Institute of Technology Madras, Chennai 600036, India
b
Formerly Shell Technology Center Bangalore, Shell India Markets Private Limited, Bangalore 562149, India
c
Department of Chemical Engineering, National Institute of Technology Hamirpur, Hamirpur 177005, India

ARTICLE INFO ABSTRACT

Keywords: Computational Fluid Dynamics (CFD) has gained a lot of popularity in the consequence assessments of gas
Gas explosion explosions in the past couple of decades. This work reviews Porosity/ Distributed Resistance (PDR) approach-
Flame acceleration based CFD solver (PDRFoam) for gas explosions. The PDR approach is used to model the effect of small-scale
CFD modelling
obstacles and only solve for the large scale congestion. PDR-based modified Favre averaged equations for mass,
momentum, enthalpy, and regress variable are solved using PDRFoam solver, which is developed as a new
application in OpenFOAM. The evaluation of the solver is done against medium-scale standard benchmark
experiments. Further, the effect of block-age ratio, different types of fuels, and different obstacle diameters
is explored. Using PDRFoam simulations, the pressure–time series, the flame arrival times, and the flame
speeds are predicted and the same are compared with experiments. Compared to standard body-fitted grid
simulations, the PDR approach is computationally economical; The predicted results are in good agreement
with experiments.

1. Introduction these studies can help to evaluate gas explosion solvers and models.
For example, experimental studies based on gas explosions are done
Accidents involving gas explosions cause significant damage to any to quantify the effect of flame propagation (Arntzen, 1998; Rogers,
gas storage or production plants. Examples of major accidents include 1995). The effect of confinement and congestion of the geometry
Flixborough (1 June 1974), Piper Alpha (6 July 1988), Buncefield (11 is investigated by modifying the bundles of obstacles in a confined
December 2005), Deepwater Horizon (20 April 2010), and Fukushima framework (Vianna and Cant, 2010; Bjerketvedt et al., 1997; Skjold
Daiichi (11 March 2011). The accidental release of gases has become et al., 2013). This is attributed to the disparate length scales of the
a repeated occurrence in the petroleum industry (Marsh, 2014). In the obstacles and establishment in the industry. Hisken et al. (2016) have
case of accidental release, a large volume of hydrocarbon gas (fuel) reported experimentally the impact of suppressing vortex shedding on
mixes with air (oxidizer) to form a flammable hydrocarbon-air mixture. the gas explosion overpressure generation. It was observed that the
If these flammable mixtures get ignited (due to any ignition source), vortex suppression can reduce the maximum overpressure by 20%–25%
an explosion might happen, leading to severe consequences in terms of in small-scale combustion.
injuries, fatalities, business interruptions, and asset damage. However, conducting experiments in large scales like a whole pro-
Quantitative Risk Analysis (QRA) studies are typically used to esti- cess plant requires a large area along with concerns about safety issues
mate the risks associated with the operation of an engineering process.
while dealing with hazardous gases. Moreover, the cost of conducting
It is used to support the understanding of the risk exposure to the em-
experiments for such scenarios is also huge. Therefore, many prac-
ployees, the environment, the company assets, and its reputation. The
tical problems consider empirical or CFD models for the estimation
risk assessment also helps to make cost-effective decisions and manages
of overpressure generation. The performance of empirical relations
the risks for the entire asset life cycle. Considering the consequences of
has not been consistent for all types of geometries and fuels (Lak-
such gas leaks, industries tend to do a lot of QRA studies. One of the
shmipathy et al., 2019). In the present investigation, the PDRFoam
most critical consequences of gas release is ignition of fuel–air mixture
tool (an open-source variant of OpenFOAM) is evaluated against a
leading to explosion. Tools are required to estimate the overpressure
database of experimental results. Different small-scale and medium-
generated during such explosions. Consequently, many experimental
studies are performed to understand the physics behind gas explosions; scale experimental investigations are used for the validation of the

∗ Corresponding author.
E-mail address: vagesh@iitm.ac.in (V.D. Narasimhamurthy).
1
Equal contribution.

https://doi.org/10.1016/j.jlp.2023.105164
Received 24 February 2023; Received in revised form 1 June 2023; Accepted 27 August 2023
Available online 30 August 2023
0950-4230/© 2023 Elsevier Ltd. All rights reserved.
M. Dhiman et al. Journal of Loss Prevention in the Process Industries 85 (2023) 105164

solver PDRFoam. The paper is presented as follows: a brief overview


of the computational methodology used in PDRFoam is described in ⎧2.0 𝐼 < 0.01
⎪ ( )
Section 2. Section 3 provides the results and discussion, comprising 𝐶𝐷 = ⎨2.0 − 0.5 𝐼−0.01
0.015
0.01 < 𝐼 < 0.025 (6)
the assessment of the performance of a tool PDRFoam against exper- ⎪
⎩1.5 𝐼 > 0.025
imental test rigs. Finally, conclusions drawn from the current work are
discussed in Section 4.
𝜕 𝜕 𝜕 𝜕̃
ℎ 𝜕
(𝛽 𝜌̄̃
ℎ) + (𝛽 𝜌̄̃
ℎ𝑢̃ 𝑗 ) = (𝐷 𝛽 )− (𝛽 𝜌̄𝑢̃
′′ ℎ′′ )
2. Computational methodology 𝜕𝑡 𝑣 𝜕𝑥𝑗 𝑗 𝜕𝑥𝑗 ℎ,𝐿 𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗 𝑗 𝑗
[ ]
𝜕 𝑝̄ 𝜕 𝜌̄𝐾̃ 𝜕 𝜌̃ ̄ 𝑢𝑗 𝐾̃
Porosity/Distributed Resistance (PDR) method has its origin since + 𝛽𝑣 − − (7)
𝜕𝑡 𝜕𝑡 𝜕𝑥𝑗
1974 when Patankar and Spalding (1974) proposed it for simulating
shell-tube type heat exchangers. Later the extension of this methodol-
ogy for problems with turbulence was reported by Sha and Launder 𝜕 𝜕 𝜕 𝜕̃
ℎ𝑢 𝜕
(𝛽 𝜌̄̃
ℎ )+ (𝛽 𝜌̄̃ 𝑢 )=
ℎ ̃ (𝐷 𝛽 )− (𝛽 𝜌̄𝑢̃
′′ ℎ′′ )
(1979) and Sha et al. (1982). Hjertager et al. (1996) used PDR for sim- 𝜕𝑡 𝑣 𝑢 𝜕𝑥𝑗 𝑗 𝑢 𝑗 𝜕𝑥𝑗 ℎ𝑢 ,𝐿 𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗 𝑗 𝑗 𝑢
[ ]
ulating gas explosions (deflagrations) in a commercial CFD code called 𝜌̄ 𝜕 𝑝̄ 𝜕 𝜌̄𝐾 ̃ 𝜕 𝜌̃ ̃
̄𝑢𝑗 𝐾
FLACS (FLame ACceleration Simulator). Further, some CFD codes using + 𝛽𝑣 − − (8)
𝜌𝑢 𝜕𝑡 𝜕𝑡 𝜕𝑥𝑗
PDR philosophy started appearing in the 1990s. EXSIM (developed
by Telemark Technological R&D Centre and Shell Global Solutions), In Eq. (8), subscript ‘u’ is used to represent variables for unburnt
AutoReaGas (developed by Netherlands Organisation for Applied Sci- gas (used in Eqs. (7) and (8)). ̃ℎ represents the enthalpy of gas. 𝐷ℎ,𝐿
entific Research (TNO) and Century Dynamics that later merged with represents the thermal diffusivity of gas. 𝑢̃
′′ ℎ′′ represents the turbulent
𝑗
ANSYS), NEWT (developed by the Cambridge University with modified ̃
enthalpy fluxes, 𝐾 is kinetic energy per unit mass.
PDR model) are few popular codes known in CFD community that can
simulate the gas explosions (Narasimhamurthy et al., 2015). 2.2. Turbulence modelling

2.1. Governing equations and PDR terms The standard 𝜅 − 𝜖 turbulence model is used with modified PDR
terms. Eqs. (9) and (10) are modified transport equations for turbulent
Eqs. (1), (2), (7), and (8) represent the PDR-based modified gov- kinetic energy (𝜅) and turbulent dissipation rate (𝜖) respectively.
erning (Favre averaged) equations for mass, momentum, enthalpy (un- [ ( ) ]
burnt and burnt) conservation and regress variable. The standard 𝜅 − 𝜖 𝜕 𝜕 𝜕 𝜇 𝜕̃
𝑘
̃
(𝛽𝑣 𝜌̄𝑘)+ (𝛽𝑗 𝜌̄̃
𝑘̃𝑢𝑗 ) = 𝛽𝑗 𝜇 + 𝑡 ̄𝜖 (9)
+(𝛽𝑣 𝑃 +𝑃𝑅 )−𝛽𝑣 𝜌̃
model (Launder and Sharma, 1974) is used along with PDR-based 𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜎𝑘 𝜕𝑥𝑗
modifications for turbulence modelling. Using the classical gradient [ ( ) ]
hypothesis, turbulent terms in Eqs. (7) and (8) are modelled. 𝜕 𝜕 𝜕 𝜇 𝜕̃
𝜖 𝜖̃ 𝜖̃2
(𝛽𝑣 𝜌̃
̄𝜖 ) + (𝛽𝑗 𝜌̃ 𝑢𝑗 ) =
̄𝜖 ̃ 𝛽𝑗 𝜇 + 𝑡 + 𝐶1 𝛽𝑣 𝑃 − 𝛽𝑣 𝐶2 𝜌̄
𝜕 𝜕 𝑚̄ 𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜎𝜖 𝜕𝑥𝑗 ̃
𝑘 ̃
𝑘
(𝛽 𝜌)
̄ + (𝛽 𝜌̃
̄𝑢 ) = (1)
𝜕𝑡 𝑣 𝜕𝑥𝑖 𝑗 𝑖 𝑉 𝜖̃
+ 𝐶1 𝑃𝑅 (10)
̃
𝑘
𝜕 𝜕 𝜕 𝑝̄ 𝜕
(𝛽 𝜌̄𝑢̃ ) + (𝛽 𝜌̄𝑢̃ 𝑢̃ ) = −𝛽𝑣 𝛿 + (𝛽 𝜏̃ ) + 𝛽𝑣 𝜌𝑔
̄ 𝑖 Here, 𝜇 is dynamic viscosity, 𝜇𝑡 is eddy viscosity, 𝜎𝑘 , 𝜎𝜖 , 𝐶1 and 𝐶2
𝜕𝑡 𝑣 𝑖 𝜕𝑥𝑗 𝑗 𝑖 𝑗 𝜕𝑥𝑖 𝑖𝑗 𝜕𝑥𝑖 𝑗 𝑖𝑗 are model constants. P is the turbulent kinetic energy production rate
𝜕
− 𝑅𝑖,𝑜 − (𝛽 𝜌̄𝑢̃′′ 𝑢′′ ) (2) term, and 𝑃𝑅 is the turbulence rate generated by sub-grid obstacles.
𝜕𝑥𝑗 𝑗 𝑗 𝑖 𝑃𝑅 is an extra source term from the PDR method which is empirically
Here, all the Favre-averaged terms are represented with an overhead modelled and depends on the length scale of obstacles (Puttock et al.,
tilde (for example ̃ 𝑢 is Favre averaged velocity). 𝜌 is the density, 𝑢𝑖 2022). All other constants in the model are taken with their standard
is the fluid velocity in 𝑖th direction, 𝜏𝑖𝑗 is stress tensor, 𝑝 is pressure, values (Launder and Sharma, 1974).
𝑔𝑖 is gravity vector and 𝑢′′ ′′
𝑖 𝑢𝑗 is the Reynolds stress tensor. The source
term for mass is represented as 𝑉𝑚̄ . The terms 𝛽𝑣 and 𝛽𝑗 represent the 2.3. Combustion modelling
volume porosity and area porosity of a cell which are PDR fields. The
term 𝑅𝑖,𝑜 is the drag produced by sub-grid obstacles, which is modelled Combustion is modelled using a laminar flamelet approach (Poinsot
by Eq. (3). and Veynante, 2005) as in most of the premixed cases. This is based on
1 the assumption that the probability of finding the reaction occurring
𝑅𝑖,𝑜 = ̄ 𝐴𝑢 |𝑢 |
𝜌𝐶 (3)
2 𝐷 𝑖 | 𝑖| is much less than the probability of finding burnt or unburnt gas.
Here, A is the wetted surface area of an obstacle and 𝐶𝐷 is the drag This implies that the flame thickness is very small as all the chemical
coefficient. Calculating 𝐶𝐷 for the approximation/modelling of drag reactions happen in the flame. Also, the smallest turbulent length scale
induced by sub-grid obstacles is an important step in the PDR approach. or eddy size is larger than the flame thickness. In such a scenario,
The current model assumes that the flow velocity (U) is uniform across combustion progress is modelled in terms of progress variable (c) or
the grid. Local Reynold’s number (Re) is calculated using the length regress variable (b). Eq. (11) represents the mathematical definition of
the progress variable. The Favre averaged regress variable (𝑏)̃ is used
scale of obstacle (D), i.e. 𝑅𝑒 = 𝑈𝜈𝐷 . Using the local 𝑅𝑒 and the relations
given by Puttock et al. (2000) are used for 𝐶𝐷 calculation. Eq. (4) is in combustion modelling, and Eq. (12) represents governing equation
used for cylinders. For the value of 𝐿𝑜𝑔10 (𝑅𝑒) in between 5.3 to 5.5, for the regress variable.
linear interpolation is used. 𝑇 − 𝑇𝑢 𝜌 − 𝜌𝑢
{ 𝑐= =1−𝑏= (11)
𝑇𝑏 − 𝑇𝑢 𝜌𝑏 − 𝜌𝑢
1.0 𝐿𝑜𝑔10 (𝑅𝑒) < 5.3
𝐶𝐷 = (4) Here 𝑇𝑢 and 𝑇𝑏 represent the temperature of unburnt & burnt gas. 𝜌𝑢
0.7 𝐿𝑜𝑔10 (𝑅𝑒) > 5.5
and 𝜌𝑏 represent the density of unburnt & burnt gas respectively.
The relations for sharp-edged obstacles have a similar form (Eq. (6)).
( ) ( )
Turbulent intensity (Eq. (5)) is also used in those relations. 𝜕 𝜕 𝜕 𝜕̃
𝑏 𝜕
𝛽 𝜌̄̃
𝑏+ ̄𝑢 ̃
𝜌̃ 𝑏= 𝐷𝑙 + 𝜌̄𝑢̃ ̃ ̃
′′ 𝑏′′ +𝛽 𝜌 𝑆 𝛯|𝛥𝑏|+𝛽𝑣 𝜌𝑢 𝑆𝐼
𝑈′ 𝜕𝑡 𝑣 𝜕𝑥𝑗 𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑗 𝑣 𝑢 𝑙

𝑈
𝐼= (5) (12)
𝐿𝑜𝑔10 (𝑅𝑒)

2
M. Dhiman et al. Journal of Loss Prevention in the Process Industries 85 (2023) 105164

Fig. 1. The geometry of nine pipe case experiments by CMR.

where, 𝐷𝑙 is the laminar diffusion coefficient, 𝑆𝑙 is the laminar flame


speed, and 𝛯 is the flame wrinkling parameter. The second term on
the right-hand side of Eq. (12) is the turbulent flux. It is modelled
using the classical gradient approach. The empirical relations are used
to calculate the laminar flame speed. The flame wrinkling (𝛯) plays an
important role in combustion simulations. Further, this term in PDR
formulation is decomposed into 𝛯𝑡 , and 𝛯𝑝 i.e. flame wrinkling due
to turbulence model and flame wrinkling due to sub-grid obstacles
respectively. Further, turbulent flame speed (𝑆𝑡 ) can be modelled as
shown in Eq. (13).
̃𝑡 𝛯
𝑆𝑡 = 𝛯 ̃𝑝 𝑆𝑙 (13) Fig. 2. (a) Flame propagation at four instants of time for the medium grid (𝛥𝑋∕𝐷 =
0.5); flame surface close to ignition (t = 0.06 s), flame reaching the first row of the
Ignition modelling is another important aspect of simulating gas obstacle (t = 0.21 s), flame reaching the second row of obstacles (t = 0.24 s) and
explosions. This is done by forcing a change in the regress variable finally flame passing through all the obstacles (t = 0.26 s), (b) flame surface depicted
by iso-surface of regress variable, 𝑏̃ = 0.5 for two different grids (c) The time trace of
directly in the cells near the ignition point. The grid is well-refined overpressure (maximum value in the entire domain) for the three different grids.
near the ignition point so that the forced change does not lead to
any numerical instabilities. For further details about all the models
and terms involved in the equations see Puttock et al. (2022). All
different grids, coarse ( 𝛥𝑋 = 1), medium ( 𝛥𝑋 = 0.5), and fine ( 𝛥𝑋 =
the equations are discretized using the finite volume method (FVM). 𝐷 𝐷 𝐷
First order Euler backward scheme is used for time advancements, the 0.25) respectively. Here, 𝛥𝑋 is the width of the computational cell in
second-order central differencing is used for all diffusion terms and a uniform mesh (𝛥𝑋 = 𝛥𝑌 = 𝛥𝑍). Numerically, ignition is carried out
the Gaussian upwind scheme is used for convective terms. The PIMPLE by varying the regress variable near the ignition point. Further details
algorithm is used for pressure–velocity coupling. Further, the pre-tuned about the ignition procedure are provided in the study by Puttock et al.
conjugate gradient method is used for solving the set of linear equations (2022).
using the OpenFOAM environment (Dhiman et al., 2022; Weller et al., As the flame propagates, it pushes the unburnt gases to the conges-
1998). tion created by 9 pipes. Fig. 2(a) and (b) show different phases in a
deflagration phenomenon. Ignition, hemispherical flame development,
3. Results and discussion
quasi-laminar flame propagation, and flame acceleration near obstacles
can be observed at different instants in time. While the flame goes
Various experiments available in the open literature are used to
through various stages of acceleration, it generates high-velocity mag-
evaluate this novel solver. First, the CMR (Christian Michelsen Re-
search) (Arntzen, 1998) experiments consisting of 9 pipe bundle is nitude flow fields (see Fig. 3). This induced flow field and the flame act
used for evaluation. Second, British Gas experiments (Rogers, 1995) are in positive acceleration feedback mode leading to turbulent flames.
considered. The simulations are performed with 20, 40, 56, and 80 pipe Flame surface area increases as the flame moves towards obstacles.
geometries with varying pitch ratios. Third, simulations are carried out The increase in the surface area is due to: (I) the wrinkling of the flame
for the SOLVEX (Shell Offshore Large Vented EXplosion) experimental by the turbulent eddies generated in the wake of the obstacles and (II)
series. Finally, the evaluation against the 3D Sotra experiments (Bjer- the stretching of flame in the venting direction. In Fig. 4, the time
ketvedt et al., 1997) is considered. trace of approximated (using regress variable 𝑏̃ = 0.5) flame surface
area and overpressure are compared. The maximum combustion rate
3.1. Nine pipe bundle-CMR leads to the maximum overpressure. We observe the close occurrence
of overpressure and flame surface area maxima (Fig. 4). Note that,
A test case with nine pipe bundle inside a tunnel as shown in the combustion rate is proportional to the flame surface area, and
Fig. 1 (Arntzen, 1998) is considered for simulations. The dimensions
PDRFoam captures this phenomenon. However, PDR and RANS-based
of the computational geometry are 3 m in length, 1 m in breadth, and
codes like PDRFoam can only give a first-order approximation of the
1 m in height respectively. The pitch (L) i.e., the distance between
flame surface.
two cylinders, is 0.33 m vertically and horizontally, and the obstacle
diameter (D) is 0.01 m. Methane-air mixture is filled in the domain Fig. 2(c) shows a comparison of pressure–time history predicted by
with an equivalence ratio of unity. The mixture is ignited at the centre the coarse, medium, and fine grids. For the case of a medium grid,
of a front wall (see Fig. 1) so that the flame accelerates towards the the predicted value of maximum overpressure (1288 mbar) is in good
outlet (venting) while passing through the congestion. agreement with the experimental value (1272 mbar). Fig. 2(c) also
For modelling, a no-slip boundary condition for flow velocity is depicts the grid sensitivity of the PDR method. Because of its grid-
specified at the front, top, side, and bottom walls of the tunnel. The sensitive nature, the PDR method should be used with the best grid
flame propagation and the pressure dynamics are investigated for three guidelines.

3
M. Dhiman et al. Journal of Loss Prevention in the Process Industries 85 (2023) 105164

Fig. 3. Velocity magnitude contours and the corresponding streamlines : (a, e) flame in laminar regime, (b, f) flame interacts with the first row of obstacles, (c, g) flame interacts
with the second row of obstacles and is split into two halves, (d, h) flame interacts with the last row of obstacles.

Fig. 5. (a) Computational geometry (length - 9 m, width - 4.5 m, and height - 4.5 m)
and the obstacle arrangement for 20 pipes case. P1 is the ignition point. (b) pitch for
Fig. 4. Comparison of overpressure and flame surface area evolution with time. 20 pipes case, (c) pitch for 40 pipes case, (d) pitch for 80 pipes case, and (e) pitch for
56 pipes case.

3.2. British gas experiments


56 pipe case, the alternate columns are repeated. More details about
The experimental cases with varying numbers of pipes (and pitch) the geometry are reported in Rogers (1995). This experimental study
inside a cuboid investigated by Rogers (1995) are considered for eval- highlights the effect of congestion on maximum overpressure, as the
uation of PDRFoam. This case, like the nine-pipe case (in Section 3.1),
overall volume of the container, remains the same, but the volume
is also ignited at the centre of the front wall. The flame then acceler-
of congestion (i.e. obstacles) varies in different experiments. Methane-
ates towards the venting direction (outlet) after it passes through the
congestion. Fig. 5(a) is the schematic of geometry in the British gas air mixture with an equivalence ratio of 1 is considered in all the
experiments. The number of columns for the 20 pipe case is four and experiments. In this case, simulations are performed using two grids,
for all other cases, it is eight. The details of pitch are shown in Fig. 5(b) coarse ( 𝛥𝑋
𝐷
= 1.66) and fine ( 𝛥𝑋𝐷
= 0.83) respectively (In the current
𝛥𝑋
for 20 pipes, (c) for 40 pipes, (d) for 80 pipes, and (e) for 56 pipes. For discussion, 𝐷 = 0.83 i.e. fine grid results are discussed).

4
M. Dhiman et al. Journal of Loss Prevention in the Process Industries 85 (2023) 105164

Fig. 8. The comparison of the maximum overpressure obtained from the British gas
experiments with the PDRFoam predictions. The solid black line represents prediction
Fig. 6. Time trace of overpressure for 20, 40, 56, and 80 pipes. matching the experiments (while the two other dashed lines represent a factor of two
under- and over-prediction).

effect on turbulence generation and the subsequent acceleration of the


flame. This acceleration of flame leads to rapid change in the mode
of flame propagation. In addition to the region outside the congestion,
the flame also moves in the region between the obstacles. This may
also cause some flame acceleration. In order to have good agreement
in the overpressure predicted by the solver and observed values in
the experiments, the solver must address the flame acceleration and
flame wrinkling. Fig. 7 shows that flame folds and merges back after
its interaction with the obstacles, i.e. on the rear side of the obstacles.
The flame folding and other modes of flame propagation are captured
by PDRFoam in all the test cases.
The flame development occurs initially in a laminar state. Due to
intrinsic flame instabilities, the freely expanding flame undergoes a
quasi-laminar state of expansion (Poinsot and Veynante, 2005). To
capture this, a quasi-laminar flame expansion model (Puttock et al.,
2022) is used in PDRFoam. As the flame accelerates, the flame area
further increases. The enhanced flame area pushes the flame front into
the unburnt gases. These unburnt gases enter into the recirculation
zones formed past the obstacles. In these recirculation zones, unburnt
gases can get accumulated. When the flame reaches these zones, more
unburnt gases are available for combustion and hence there is a sudden
rise in the combustion rate. This sudden increase in the combustion rate
can be related to the sudden pressure rise. Hence, one of the reasons
behind the pressure peaks can be an accumulation of unburnt gases in
the wake zone of obstacles. This hypothesis is based on the fact that
the time instants of the merging of the flame in the wake of obstacles
and the occurrence of maximum overpressure are qualitatively similar.
Fig. 8 shows the quantitative analysis of predicted maximum over-
pressure using simulations. The horizontal axis shows the experimental
overpressure values while the vertical axis depicts the simulated over-
pressure values and the under-prediction and over-prediction by a
factor of two are shown as dashed lines in Fig. 8. It is evident that the
simulation results are in good agreement with the experimental find-
Fig. 7. Flame surface (fine grid) depicted by iso-surface of regress variable, b = 0.5; ings. However, slight over-prediction is observed for higher congestion
for various cases with a different number of pipes at different time instants. cases.
The effect of obstacle spacing and congestion on the maximum
overpressure can be seen in Fig. 9. The under-prediction is observed for
Fig. 6 depicts the pressure–time history for different congestions lower congestions (i.e. 20 and 40 pipes). However, for higher conges-
i.e., varying the number of pipes as mentioned earlier. The black colour tions (56 and 80 pipes) the solver can predict maximum overpressure
profile corresponds to the 80 pipe case and the maximum overpres- with certain positive deviations i.e., for higher congestions, the solver
sure value is seen in that case. One of the important reasons behind over-predicts maximum overpressure. It was reported that when the
overpressure peaks is flame acceleration. The presence of obstacles same experiments are repeated, there was an uncertainty of about 25%
(pipe bundles here) and the walls of the tunnel has an immediate in the maximum overpressure recorded (Puttock et al., 1996). Such

5
M. Dhiman et al. Journal of Loss Prevention in the Process Industries 85 (2023) 105164

than that of methane; also heat release of propane during combustion


is more in comparison to methane. This means that the burning of
an equal volume of propane will release more energy compared to
methane, which may lead to a higher magnitude of overpressure and
flame speeds.

3.3.3. Effect of obstacle position


The Fig. 12 compares the effect of the position of obstacles. In
Fig. 12(a) methane is used as fuel and in 12(b) propane is used. In
Fig. 12(a), the maximum overpressure is observed for two column
cases. This is in good agreement with the experimental findings. An
increase in the number of obstacles i.e., higher congestion produces
higher flame speeds leading to a higher combustion rate and hence it
leads to the highest maximum overpressure value. This trend has been
captured in the British gas cases also (see Section 3.2). The occurrence
Fig. 9. Effect of the number of obstacles on the maximum overpressure in all the of maximum overpressure (the time instant at which it occurs) is
British gas test rigs. different depending on the obstacle position. For the front column and
two-column configurations, maximum overpressure is seen earlier than
the rear column configuration.
experimental uncertainties could be due to pressure sensor malfunction, In Fig. 12(b), i.e. in propane cases, significant over-prediction is
probe location influence, etc. Note that most experimental evidence seen in the rear column case compared to experimental observation.
in the field suggests an increase in overpressure with an increase in The reason for the same is being explored. Sensitivity towards the PDR
congestion, i.e., similar to the current simulations. approach can be one of the reasons for this. To analyse the sensitivity
towards the PDR approach and the domain size, we have simulated
3.3. SOLVEX cases the rear column propane case with different grid resolution. Also,
we extended the domain size by five times the actual cloud size to
3.3.1. Geometry analyse sensitivity towards domain size (boundary condition effects).
SOLVEX (Arntzen, 1998) experimental test series is also considered This analysis is presented in an Appendix. It was found that the SOLVEX
for evaluating the solver PDRFoam. Fig. 10 is the schematic diagram for rear column propane case is highly sensitive to domain size (boundary
different SOLVEX cases. Three different geometric configurations were conditions) and grid size. Please see the appendix section (Appendix C)
investigated experimentally along with two different fuels. Different ge- for more details.
ometric configurations are: (I) front column (II) dual column, and (III) The Fig. 14 shows the predicted maximum overpressure values
rear column. For understanding the effect of different fuels, methane where PDRFoam results are in good agreement with the experiments.
and propane are used. Fig. 10(a) shows the configuration having two The solid 45◦ line (black line) represents the ideal case of peak
columns of obstacles (schematic of top view), Fig. 10(b) shows the front predicted maximum pressure equal to the experimental value. The
column and Fig. 10(c) shows the rear column configuration. Vianna other two lines (dashed lines) are used to denote the under and over-
and Cant (2010) have discussed details of the placement of pipes (for prediction of maximum overpressure by a factor of two. Front column
further details of geometry). configurations of methane and propane, and rear column configura-
tion of propane are over-predicted. The other three configurations
3.3.2. Effect of different fuels are under-predicted but overall there is good agreement with the
Fig. 11(a), (b), and (c) show the comparison for pressure–time experiments.
history predicted using PDRFoam for all configurations i.e., two, front,
and rear columns respectively, with methane (grey colour) and propane 3.4. 3D-Sotra experiments
(black colour) as fuel. For the two-column configuration, irrespective of
fuel type two pressure peaks are observed. This is due to the presence The evaluation of the PDRFoam is also extended to the 3D corner
of two separate columns of obstacles and the gap available between the Sotra experiments conducted by British Gas (Bjerketvedt et al., 1997).
two columns. There is an increase in flame surface area due to the first The 27 m3 cornered cube consists of pipes, where the number and
set of obstacles. Afterward, the flame merges back and again due to the size(s) are varied from 2 × 2, 3 × 3, and 4 × 4. Two different pipe
second set of obstacles the flame expands. The merging of flame after it diameters are used 0.41 m and 0.82 m. Fig. 15(a) is the front and (b)
has crossed the first set of obstacles is possible because of the sufficient is the plan view of the cubical vessel and 3 × 3 pipes. The three sides
gap between the columns. This increases the burning rate or rate of of the cubical vessel are of polythene (used to confine the cloud size)
combustion product formation twice during the flame propagation. denoted by dashed lines in the schematic. The propane-air mixture is
Also, it leads to the interaction of flame with accumulated unburnt ignited at one corner of the wall (see Fig. 15) having an equivalence
gases twice. This may be one of the reasons for the two peaks in the ratio of unity. The flame propagation at various instants of time can
overpressure. Similar trends are seen in a study conducted by Vianna be visualized in Fig. 16. The flame front gets distorted and extended
and Cant (2010). In all configurations, it is evident that maximum to various regions of the cube after interacting with the obstacles. The
overpressure is more in the case of propane compared to methane. This flame speed increases substantially due to the congestion in the domain.
observation is in good agreement with experimental investigations. The available literature lacks the data for flame speed; it is impossible
Fig. 11(d), (e), and (f) depicts the flame arrival at different locations. to compare the flame predictions with experiments.
The slope of the flame arrival curve depicts the flame speed at different
instants in time. It was found that during the instants at which flame 3.4.1. Effect of Volume Blockage Ratio (VBR) and increase in number of
interacts with obstacles (cf. Fig. 13), the propane flame arrival curve obstacles
(black curve) has a greater slope than the methane (grey curve). The effect of the Volume Blockage Ratio (VBR) on the over-pressure
Clearly, the flame speed for propane deflagration is more than methane. is also investigated. Fig. 17(a) shows the overpressure variation for two
This can be attributed to the reactivity of propane which is higher VBRs. The higher congestion configuration (i.e. VBR = 0.2) gives the

6
M. Dhiman et al. Journal of Loss Prevention in the Process Industries 85 (2023) 105164

Fig. 10. Computational domain (length - 10 m, width - 6.25 m, and height - 8.75 m) and the obstacle geometry (diameter - 0.5 m and length - 8.75 m) for SOLVEX pipe bundle
case. P1 is the ignition point, which is the same for cases in (a), (b), and (c). (a) is two columns, (b) is the front column, and (c) is the rear column configuration.

Fig. 11. Time trace of overpressure in (a) two-column case, (b) front column case, and (c) rear column case. Flame arrival at different locations in vent direction in (d) two-column
case, (e) front column case, and (f) rear column case.

maximum overpressure significantly more compared to the lower con- 1997). PDRFoam can capture this trend qualitatively. For the same
gestion configuration. This also envisaged the fact that gases accumu- VBR, i.e. 0.2, we observe higher overpressure value for 4 × 4 configu-
late around the obstacles. So, more blockage means more accumulation ration (i.e. 16 pipes) compared to 2 × 2 configuration (i.e. four pipes)
of gases and hence, higher maximum overpressure values. Fig. 17(b) (Fig. 17(b)). Fig. 18 shows the comparison of maximum overpressure
presents the test cases for Sotra experiments at VBR = 0.2. The maxi- for all the test cases done for 3D Sotra experiments. A significant
mum overpressure for two different obstacle sizes (0.41 m, and 0.82 m) over-prediction can be seen for VBR = 0.5 with the experimental
is obtained in 17(b). The effect of obstacle size in the same VBR observations. This indicates that the PDR method is sensitive to VBR
(i.e. 0.2) is evaluated. In the literature (Bjerketvedt et al., 1997), for the as seen earlier also in the case of British gas simulations (Section 3.2).
same volume blockage ratio (VBR), the highest overpressure value is For higher congestion (or VBR), the PDRFoam generally shows over-
observed for the smallest diameter case. This can be because more shear prediction. Further investigation is required in the context of grid
layers are formed for smaller diameter geometry (Bjerketvedt et al., generation (and grid guidelines are required) for various geometries.

7
M. Dhiman et al. Journal of Loss Prevention in the Process Industries 85 (2023) 105164

Fig. 12. Time trace of overpressure for (a) methane as a fuel and (b) propane as a fuel. In both (a) and (b), black represents two columns case, blue depicts the front column
case and the green depicts the rear column case. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 13. Flame surface depicted by iso-surface of regress variable, b = 0.5; at various
time instants for the rear column case with propane-air mixture.

Fig. 15. (a) Front and (b) plan view of geometry for 3D Sotra cubical vessel with
3 × 3 pipe bundle arrangement, P1 is the ignition point. The dashed lines are used to
show sides with polythene.

3.5. Statistical analysis

Statistical analysis based on the geometric mean–variance (V), and


the mean bias (B) is also performed on the data obtained through
simulations as suggested by the MEGGE group (Snowdon et al., 1999).
The definitions for V and B are given in a recent study by Puttock et al.
(2022). The maximum overpressures for all the experiments are shown
in Fig. 19 using the geometric variance plotted against the geometric
mean. The model prediction is good with 100% accuracy if the value of
B and V is equal to 1, i.e., the vertex (1,1). The vertical line divides the
under and over-prediction into the obtained data points. The left side
data point indicates the under-prediction, and the right side suggests
the over-prediction in the obtained results. It is observed that there is
over-prediction for 9 pipe and British gas cases. However, SOLVEX and
Sotra experiments are slightly under-predicted.
Fig. 14. The comparison of the maximum overpressure obtained from the SOLVEX
experiments with the PDRFoam predictions. The solid black line represents prediction
matching the experiments (while two other dashed lines represent a factor of two
4. Conclusion
under- and over-predictions).
In the current study, the PDR approach to model the effect of
small-scale obstacles, e.g., pipes on flame propagation and explosion
overpressure, is investigated. The PDR approach ascertains the peak

8
M. Dhiman et al. Journal of Loss Prevention in the Process Industries 85 (2023) 105164

Fig. 16. The flame surface propagation at different instants of time corresponding to iso-surface of regress variable b = 0.5 for the 3 × 3, Volume Blockage Ratio (VBR) of 0.1
cases.

the flame propagation is well addressed. The flame wrinkling effect is


also observed and is pronounced at higher refinements. The predicted
peak pressure appears to be in close agreement with the experimental
observations. However, the predicted peak pressure does vary with grid
refinements. Thereby suggesting the vulnerability of the PDR method
with grids and hence demanding best practice guidelines. The mesh
count in higher refinement cases here is significantly low compared to
a mesh count in a standard body-fitted approach. PDRFoam indicates
over-predictions for peak pressure in most cases indicating its conser-
vative nature. Therefore, it can be stated that the solver PDRFoam is
computationally cost-effective and can be a good alternative to the stan-
dard body-fitted approach for explosion modelling using CFD. Further,
Fig. 17. (a) The time trace of overpressure for two VBRs 0.1 (black) and 0.2 (grey)
(b) The time trace of overpressure for two obstacle sizes for the same VBR (i.e. 0.2): the very nature of open-source code provides many more modelling
4 × 4 (black) and 2 × 2 (grey). opportunities.

CRediT authorship contribution statement

Manish Dhiman: Methodology, Software, Validation, Formal anal-


ysis, Writing – original draft. Anand Zambare: Methodology, Software,
Validation, Formal analysis, Investigation, Writing – original draft.
Pratap Sathiah: Conceptualization, Methodology, Software, Investiga-
tion, Writing – review & editing. Vagesh D. Narasimhamurthy: Con-
ceptualization, Methodology, Formal analysis, Investigation, Resources,
Writing – review & editing, Supervision, Funding acquisition.

Declaration of competing interest

The authors declare that they have no known competing finan-


cial interests or personal relationships that could have appeared to
Fig. 18. The comparison of the maximum overpressure obtained from the 3D Sotra influence the work reported in this paper.
experiments with the PDRFoam predictions. The solid black line represents prediction
matching the experiments (while two other dashed lines represent a factor of two
under- and over-prediction). Data availability

Data will be made available on request.


overpressure and the regress variable statistics for the combustion test
cases comprising tunnels with congestion of obstacles. The test cases
comprising CMR, British Gas, SOLVEX, and Sotra experiments having
Acknowledgements
methane-air mixture in the domain are explored. Propane gas is also
used to simulate one test rig of SOLVEX and 3D Sotra test cases.
The explosion parameters such as flame speed, flame arrival time, This work is partially supported by the Data Analytics, Risk and
peak pressure, and regress variable contours are obtained to assess the Technology (DART) laboratory, American Express India Private Lim-
validity of the solver. A comparison of the regress variable contour ited. We also acknowledge the P.G. Senapathy Center for Computing
plots and flame propagation images at different instances suggest that Resource, IIT Madras for a grant of computing time.

9
M. Dhiman et al. Journal of Loss Prevention in the Process Industries 85 (2023) 105164

Fig. 19. The comparison of the geometric mean–variance (simulation/experiment) against geometric mean bias (simulation/experiment) for PDRFoam (b) The zoomed view for
the data points.

Appendix A. PDRFoam setup Table 1


Boundary conditions used in PDRFoam simulations.
Variable Wall B.C. Venting B.C.
b, ft, GRep, GRxp, zeroGradient inletOutlet
• Step I: Geometry modelling RPers, Su, T, Tu, Xi, Xp
In the current version of PDRFoam (OpenFOAM V2012), the U fixedValue (U = 0.0) inletOutlet
p zeroGradient waveTransmissive (𝛾 = 1.3)
obstacle shapes can be cylindrical or cuboids. It is possible to vary epsilon, k, nut wallFunction inletOutlet
the porosity of every obstacle. The orientation of the obstacle
is also constrained to three principal axes. Considering these
limitations, modellers can approximate the complex geometry of
practical situations. The authors recommend keeping the actual
and approximated geometry volume the same, i.e., the global
volume blockage ratio should be the same. The overall domain
size depends on the type of explosion scenario. For an internal
explosion (as in most of the studies in the current manuscript),
only obstacles can be modelled with an external domain size
equal to the container/gas cloud size. For external explosion
cases, the external environment should also be modelled. The
domain size can be considered five times the actual cloud volume
in such cases. In these case, the mixture container itself is treated
as an obstacle.

• Step II: Gas composition and location


Since PDRFoam code is used only for premixed explosions,
the user must define the premixed gas volume and composition in
the ‘SetFieldsDict’ file in the system folder of each case. The equiv-
alence ratio is considered through the mass fraction of the gas (ft).
To calculate the (ft) mass fraction, the authors have used ‘Gaseq’,
an open-source code. Modellers can use any code to calculate this Fig. 20. Flame arrival at different locations in the vent direction (Nine pipe CMR case).
mass fraction value and define that in the ‘SetFieldsDict’ option.

Appendix C. SOLVEX rear column propane case


• Step III: Ignition modelling
Ignition location, diameter, strength, and duration are used We observed an unusual secondary peak in the time trace of over-
to model ignition in PDRFoam. Ignition and combustion-related pressure for the SOLVEX rear column case with propane fuel. The
parameters are defined in the ‘constant − > combustionProper- following sensitivity studies have been performed to investigate the
ties’ file. same.
• Step IV: Boundary conditions
The (Table 1) describes the boundary conditions used for the Influence of grid size
simulations reported in the current manuscript. The Porosity Distributed Resistance (PDR) approach is grid-
dependent as mentioned earlier. In the SOLVEX case, the cylinder
diameter is the smallest geometric length scale (D). In terms of D, we
Appendix B. Flame arrival
used 0.8D, 0.4D, and 0.2D grid sizes for simulation. All of the grid sizes
show conservative predictions (Table 2). The time trace of overpressure
In this section, we provide the additional data of flame arrival time (maximum in the whole domain) for all grids show secondary peaks
from (a) the Nine pipe CMR case (Fig. 20), (b) the British gas case (Fig. 23). However, the result for grid ratio of 0.8 gives a secondary
(Fig. 21), and (c) the 3D Sotra case (Fig. 22). peak after the first pressure peak, which occurs due to interaction with

10
M. Dhiman et al. Journal of Loss Prevention in the Process Industries 85 (2023) 105164

Fig. 23. The time trace of overpressure for various grids in SOLVEX rear column
propane case.
Fig. 21. Flame arrival at different locations in the vent direction (British gas case).

Fig. 24. The time trace of overpressure in different domains for the SOLVEX rear
column propane case.
Fig. 22. Flame arrival at different locations in the vent direction (3D Sotra case).

Table 2 thereby highly sensitive to the domain size (boundary condition) and
Various grid sizes used in SOLVEX rear column propane case. the grid size. Note that, in the previous study of Puttock et al. (2022),
𝛥𝑋∕𝐷 Overpressure (mbar) PDRFoam did not show consistent behaviour in some simulations (for
0.8 1329.24 example see Figure 15, Figure 24, Figure 28 in Puttock et al. (2022)).
0.4 1417.54
0.2 1412.21

References

obstacles; thereafter the time of pressure rise in 0.8D grid case is more Arntzen, B.J., 1998. Modelling of turbulence and combustion for simulation of gas
realistic and the coarse grid gives better results for this specific case of explosions in complex geometries. (Ph.D. thesis). Norwegian University of Science
and Technology, NTNU.
SOLVEX rear column with propane as fuel.
Bjerketvedt, D., Bakke, J.R., Van Wingerden, K., 1997. Gas explosion handbook. J.
hazard. mater. 52 (1), 1–150.
Influence of domain size Dhiman, M., Meysiva, V., Sathiah, P., Narasimhamurthy, V.D., 2022. Poros-
ity/Distributed Resistance (PDR) modelling in the CFD solver PDRFoam. In: Recent
The venting geometry in the SOLVEX case is different compared to
Advances in Applied Mechanics: Lecture Notes in Mechanical Engineering. pp.
all other cases in this manuscript. To analyse the sensitivity towards 503–519.
venting boundary condition (domain size), we simulated the SOLVEX Hisken, H., Enstad, G., Narasimhamurthy, V.D., 2016. Suppression of vortex shedding
propane case with an extended domain. The domain size is taken as and its mitigation effect in gas explosions: an experimental study. J. Loss Prev.
Process Ind. 43, 242–254.
five times the actual cloud size. A grid size of 0.8D is used here.
Hjertager, B., Saeter, O., Solberg, T., 1996. Numerical modelling of gas explosions:
The secondary peaks seen in the exact domain case diminished in a review. In: International Symposium on Hazards, Prevention and Mitigation of
the extended domain case (see Fig. 24). The SOLVEX propane case is Industrial Explosion, Netharlands, Vol, 2. pp. 77–91.

11
M. Dhiman et al. Journal of Loss Prevention in the Process Industries 85 (2023) 105164

Lakshmipathy, S., Skjold, T., Hisken, H., Atanga, G., 2019. Consequence models for Puttock, J., Yardley, M., Cresswell, T., 2000. Prediction of vapour cloud explosions
vented hydrogen deflagrations: CFD vs. engineering models. Int. J. Hydrogen using the SCOPE model. Journal of Loss prevention in the Process Industries 13
Energy 44 (17), 8699–8710, Special issue on The 7th International Conference (3-5), 419–431.
on Hydrogen Safety (ICHS 2017), 11-13 September 2017, Hamburg, Germany. Rogers, M., 1995. Project EMERGE measurement of turbulent characteristics and drag
Launder, B.E., Sharma, B.I., 1974. Application of the energy-dissipation model of loads on pipework obstacles in steady unidirectional flows. Br. Gas Res. Technol.
turbulence to the calculation of flow near a spinning disc. Lett. heat mass transfer GRCR 610.
1 (2), 131–137. Sha, W., Launder, B., 1979. Model for Turbulent Momentum and Heat Transport in
Marsh, 2014. The 100 largest losses 1974–2013—Large property damage losses in the Large Rod Bundles.[LMFBR]. Technical report, Argonne National Lab., IL (USA).
hydrocarbon industry. 23rd edition, London: Marsh global energy risk engineering. Sha, W., Yang, C., Kao, T., Cho, S., 1982. Multidimensional numerical modelling of
Narasimhamurthy, V.D., Hisken, H., Atanga, G., Skjold, T., 2015. Porosity/Distributed heat exchangers. J. Heat Transfer 104 (3), 417–425.
Resistance modelling for industrial CFD applications. In: 8th National Conference Skjold, T., Pedersen, H., Bernard, L., Middha, P., Narasimhamurthy, V.D., Landvik, T.,
on Computational Mechanics (MekIT 2015), CIMNE, Trondheim, Norway. pp. Lea, T., Pesch, L., 2013. A matter of life and death: validating, qualifying and
321–331. documenting models for simulating flow-related accident scenarios in the process
Patankar, S.V., Spalding, D.B., 1974. A calculation procedure for the transient and industry. Chem. Eng. Trans. 31, 187–192.
steady-state behavior of shell-and-tube heat exchangers. Heat Exch.: Design Theory Snowdon, P., Puttock, J., Provost, E., Cresswell, T., Rowson, J., Johnson, R., Masters, A.,
Sourceb. 155–176. Bimson, S., 1999. Critical design of validation experiments for vapour cloud
Poinsot, T., Veynante, D., 2005. Theoretical and Numerical Combustion. RT Edwards, explosion assessment methods. In: Proceedings of International Conference and
Inc.. Workshop on Modeling the Consequence of Accidental Releases of Hazardous
Puttock, J., Cresswell, T., Marks, P., Samuels, B., Prothero, A., 1996. Explosion Materials, San Francisco. pp. 831–849.
assessment in confined vented geometries. SOLVEX large-scale explosion tests and Vianna, S.S., Cant, R.S., 2010. Modified porosity approach and laminar flamelet
SCOPE model development. Proj. rep. Health Saf. Exec. OTO 96 (4). modelling for advanced simulation of accidental explosions. J. Loss Prev. Process
Puttock, J., Walter, F., Chakraborty, D., Raghunath, S., Sathiah, P., 2022. Numerical Ind. 23 (1), 3–14.
simulations of gas explosion using Porosity Distributed Resistance approach part- 1: Weller, H.G., Tabor, G., Jasak, H., Fureby, C., 1998. A tensorial approach to compu-
Validation against small-scale experiments. J. Loss Prev. Process Ind. 75, 104659. tational continuum mechanics using object-oriented techniques. Comput. phys. 12
(6), 620–631.

12

You might also like